Skip to main content

Full text of "Ward identity and optical-conductivity sum rule in the d-density wave state"

See other formats


Ward identity and optical-conductivity sum rule in the d-density wave state 



L. Bcnfatto 1 ,* S.G. Sharapov 2 ,* N. Andrenacci 3 , and H. Beck 3 * 
1 Department de Physique, Universite de Fribourg, Perolles, CH-1700 Fribourg, Switzerland 
2 Istituto Nazionale per la Fisica della Materia (INFM); 
Institute for Scientific Interchange, ma Settimio Severn 65, 1-10133 Torino, Italy 
3 Institut de Physique, Universite de Neuchatel, CH-2000 Neuchatel, Switzerland 

(Dated: February 2, 2008) 

We consider the role of the Ward identity in dealing with the transport properties of an interacting 
system forming a d-wave modulated charge-density wave or staggered flux phase. In particular, we 
address this issue from the point of view of the restricted optical-conductivity sum rule. Our aim is to 
provide a controlled approximation for the current-current correlation function which allows us also 
to determine analytically the corresponding sum rule. By analyzing the role of the vertex functions 
in both the microscopic interacting model and in the effective mean-field Hamiltonian, we propose 
a non-standard low-energy sum-rule for this system. We also discuss the possible applicability of 
these results for the description of cuprate superconductors in the pseudogap regime. 

PACS numbers: 71.10.-w, 74.25.Gz, 74.72.-h 



I. INTRODUCTION 



In the last years a quite important re-examination of the optical conductivity of high-T c superconductors (HTSC) 
has been performed, due to the improved experimental resolution. Despite the variety of features observed in the 
different families of cuprates, when the integral up to large frequencies of the optical spectra is concerned a common 
behavior can be found [1-6]. This result is particularly interesting, because it would allow us to distinguish between 
different theoretical scenarios for HTSC, in particular for the pseudogap phase observed in underdoped compounds. 
The optical spectral weight is defined as the integral of the optical conductivity in a given direction i = x,y, z: 

Rea u (uj,T)du, (1.1) 

-win 

and can be analyzed as a function of both the temperature T and the cutoff frequency ujm- According to this definition, 
the weight Wi includes also the condensate peak at u = which develops in the superconducting (SC) state below 
T c . Depending on the cut-off ldm the sum rule (1.1) acquires different meanings. When all the optical transitions are 
taken into account, Eq. (1.1) expresses simply the so-called full f-sum rule [7-9], relating the optical spectral weight 
to the total carrier density n, 



r°° ■nne 2 

/ Rea(cu)duj = , (1.2) 

J-oo rn 

where m is the bare electronic mass. However, it is usually assumed that when lj^i is of the order of the plasma 
frequency only intraband optical transitions relative to the lowest conduction band £k contribute to W{T), so that 
one obtains the restricted or partial sum rule [10-12], which relates Wi to the average value of the diamagnetic term 
Tu (see Eq. (2.10) below), 



Wi(u;p,T) = W(T) = 




where n^ a is the momentum occupation number, V is the unit-cell volume, N is the number of unit cells, e is the 
electron charge, and we set H = c = 1. In the 2D case V = a 2 , and in the quasi-2D case V = a 2 s, where a is the 



'Electronic address: lara.benfatto@unifr.ch 
t Electronic address: sharapovSbitp.kiev.ua 
tElcctronic address: Hans .BeckOunine . ch 



2 



lattice spacing and s is the distance between the layers. In the following we will consider mainly in-plane processes 
and isotropic systems, where W x = W y = W . 

The main difference between the restricted and full sum rule is that while W(u>m ~> oo,T) is a constant, W(T) 
given by Eq. (1.3) is in general a function of temperature, which provides information about the interactions between 
the electrons in the system. In particular, in a 2D lattice model with a nearest-neighbors tight-binding dispersion 
ek = — 2t(cosk x a + cos k y a) the spectral weight Eq. (1.3) is proportional to the mean kinetic energy of the system, 

W(T) = —■^^p-- In the absence of interactions = /(£k), where £k = £k — fJ>, A* i s the chemical potential, and 
f(x) is the Fermi function. In this case the main temperature dependence of the spectral weight (1.3) comes from the 
temperature smearing of the Fermi function, and can be easily evaluated using the Sommcrfcld expansion: 

W(T) 1 ^ f W(0) it 



(ire 2 a 2 /V) N 

where N(e) is the density of states for the tight-binding dispersion and c(e) = eN'(e) + N(e). By making a quadratic 
approximation for the two-dimensional tight-binding band dispersion one would find c(/i) = l/Airt, which is also a 
good estimate of the exact value obtained using the true band dispersion and by doping the system away from half- 
filling (sec also Appendix A). However, for an interacting system can acquire in general a different temperature 
dependence, which influences also W(T). An example is provided by the case of a SC instability. Indeed, according 
to the BCS theory, in the SC state the occupation number becomes 



n k „ = [1 - £ k /£k tanh(^ c /2T)], (1.5) 

where Ak is the SC gap and E^ c = y£k + A 2 , is the quasiparticle dispersion in the SC state, so that the spectral 
weight (1.3) decreases below T c . When W(T) corresponds to the kinetic energy this result is understood as the 
increase of < K > below T c due to the particle-hole mixing in the SC state. 

These general expectations about the behavior of the restricted optical sum rule were not confirmed, within several 
respects, in the experiments on HTSC. Early measurements of the c-axis spectral weight up to frequencies of the 
order of the plasma edge, iop cc 10 4 cm -1 , showed that in YB 2 Cu306+<5 (YBCO) compounds W Z (T) exhibits a quite 
anomalous temperature dependence, with a decrease below the pseudogap temperature, followed by an increase below 
T c [13]. Such a behavior was indeed attributed to the effect of pseudogap opening, combined with the tunneling 
character of the transport along the c-axis direction. 

Recently more attention has been instead devoted to the issue of the spectral- weight behavior for the in-plane optical 
conductivity, which is a better probe of the degrees of freedom mostly responsible for the properties of HTSC. The 
measurements were performed in Bi 2 Sr 2 CaCu 2 8+5 (BSCCO) [1, 2, 4], YBCO [3, 5] and La 2 _ x Sr x Cu 2 4 (LSCO) [6] 
compounds at various cut-off frequencies uim between 1000 cm -1 (0.12 eV) and 20000 cm -1 (2.5 cV). A first issue 
is the behavior of W(T) below T c . While early measurements in BSCCO samples show that there is an even faster 
increase of W(T) below T c [1, 2], contrary to the prediction of the BCS theory, more recent results in BSCCO [4] 
show that there is a flattening of W(T) in underdoped samples for ujm = 8000cm -1 , while a BCS behavior below T c 
is seen in the overdoped BSCCO and in YBCO samples [3, 5]. Also from the theoretical point of view many proposals 
arose relative to the problem of the lowering of in-plane kinetic energy in the SC state [14-17]. 

Interestingly the behavior of W(T) above the SC transition also shows unexpected features, which deserve more 
investigation. Indeed, as observed in Refs. [2], the in-plane optical sum rule does not show any decrease below the 
temperature at which the pseudogap forms, contrary to what found for the c-axis response. In addition, when the 
plasma edge is considered as a cut-off, W(T) shows a "standard" T 2 temperature dependence, even though these 
are clearly strongly-interacting non-Fcrmi-liquid systems. However, this result is misleading, because despite the 
qualitative analogy with the free tight-binding result (1.4), the measured W(T) is in a strong quantitative disagreement 
with the estimate (1.4). Indeed, as we show in Appendix A, the coefficient c(/i) of Eq. (1.4) is about one order of 
magnitude larger than expected by using a t value estimated by other probes (as photoemission measurements of the 
Fermi surface), showing that the sum rule is far from being conventional already in the normal (non SC) state [6]. 
Moreover, even faster increase of W(u>m,T) is observed at smaller values of u>m [4, 6]. 

For these reasons, the issue that we address in the present paper is the behavior of the optical-conductivity spectra 
and sum rule above T c , but within a model system for the pseudogap state. Between the several proposals existing 
in the literature about the origin of the pseudogap [18], we focus in the present paper on the case where a competing 
order parameter is formed before the SC state is established. In particular, we refer to the so-called flux phase or 
(i-density wave state (DDW) [19-27]. We would like to stress that while a flux phase does not present modulated 
charge, the same phenomenological spectrum can be considered as emerging due to the tendency of the system to 
form charge order near a quantum critical point [28]. This scenario was studied in Ref. [26], and we will refer in 



3 



the present paper also to this point of view, which could be useful in relating the results presented here not only 
to cuprates, where they can be only partly applied, but also to other materials displaying a true fc-space modulated 
CDW (as for example dichalcogenide materials [29, 30]). 

In a previous publication [31], we discussed briefly how a mean- field description of the DDW state can be compatible 
with an increase of the spectral weight below the temperature at which the order parameter forms. However, this 
result was not considered from a more general point of view, which consists in relating the sum rule to the problem 
of providing a gauge-invariant approximation for the response functions in a given microscopic model. As we shall 
see, the basic requirement of respecting the charge conservation imposes simultaneously several constraints on the 
definition of the current operator, the diamagnetic term and the corresponding electromagnetic correlation functions. 
The sum rule then follows naturally when all these requirements are satisfied within a given approximation for the 
microscopic interacting model, and different approximations can lead to different sum rules. As we shall see, while 
the anomalous sum rule derived in Ref. [31] can be proposed to reproduce the experimental data for cuprates, the 
agreement with the theoretically obtained optical conductivity is more subtle, and more detailed features specific of 
different materials should be considered. A more difficult task is to properly define the change of behavior of the sum 
rule at different cut-off u>m' this problem is quite general, and while it is clear that for u>m — > oo the full sum rule 
(1.2) must be recovered, there is as yet no clear understanding of a proper experimental and theoretical definition of 
the correct cut-off for the restricted sum rule in Eq. (1.3). In our case, we shall discuss how the various restricted sum 
rules should be realized at different energy scales, even though an exact result cannot be obtained in this respect. 

The structure of the paper is the following. We begin by presenting in Sec. II the general formalism which is needed 
to analyze the optical-conductivity sum rule in an interacting system. In Sec. Ill we explicitly study the case of a 
DDW state, and we show that we can derive a good approximation for the low-energy optical conductivity which is 
however no more related to a known sum rule. In Sec. IV we solve this problem by analyzing directly the reduced, 
low-energy DDW model Hamiltonian, and we calculate explicitly the sum rule and the optical conductivity within the 
proposed mean-field approach to the DDW transition. We then discuss in Sec. V the results obtained and summarize 
the procedure described in the paper. In Appendix A we report the evaluation of the sum-rule behavior for the 
non-interacting tight-binding model, to quantify the discrepancy with the experimental data, and some details about 
the role of disorder are presented in Appendix B. 



II. SUM RULE IN A MODEL WITH GAUGE INVARIANT INTERACTION 



Let us start by considering a general Hamiltonian describing interacting electrons in a two-dimensional lattice: 

H = - t Yj C ^ C J' CT ~ ^£4^ + 4a c ™ v ( r i ~ ^Hv'^v (2.1) 



where the field operator c\ a creates an electron of spin a at r^, t is the hopping parameter, (ij) is the sum over nearest 
neighbor sites, V(ji — r^) is the translationally invariant electron-electron interaction. When rewritten in reciprocal 
space, the band dispersion corresponds to e(k) = — 2i(cos k x a + cos k y a). Throughout the paper units h = ks = c = 1 
are chosen. 

In the DDW state a particle- hole coupling is considered at the characteristic wave- vector Q = (ir/a, ir/a). The 
notation is then simplified by halving the Brillouin zone and introducing two-component electron operators (the DDW 
equivalent of Nambu spinors [32]) 



Cko- 
C k+Q,cr 



Xk 



t - J 



c 



t 

k+Q,cr 



(2.2) 



where cl and Cker arc the Fourier transforms of c\ a and c^. The Hamiltonian (2.1) written in terms of x becomes 



RBZ 



2 ( £ k + £k+q) - H + ^(ek - Ek+qKa 



k,cr 

^ BZ 



RBZ 



RBZ 



(2.3) 



N 



where V(q) is the Fourier transform of the potential, 173 is the Pauli matrix and the sums are over the reduced (RBZ) 
and full Brillouin zone (BZ). 



4 



After employing the nesting property £k+Q = — £k the kinetic term of the Hamiltonian (2.3) takes a simple form 

RBZ 

Ho = j2xL[wa-l4Xk*- (2-4) 

Accordingly, the free electron Green's function reads 

Gq 1 (k, iuj n ) = (iu n + (jb)a - £k03, (2.5) 

where w„ = (2n + 1)ttT is the fermionic (odd) Matsubara frequency. The full Green's function of the system 
G(p), p = (p, iijS) is given by the Dyson equation 

G- 1 (p) = Go 1 (p)-X(p), (2.6) 

where the self-energy £(p) is evaluated at Hartree-Fock level as 

S(p) = ^G(fe)V(p-k). (2.7) 

k 

where J^k is a short hand notation for T/NJ2i^ n J2k BZ ■ Observe that for a time-independent interaction, as the 
one considered in Eq. (2.1), the self-energy (2.7) docs not depend explicitly on the frequency, but we will keep 
for convenience this more general notation in the following. In the case of superconductivity, the Hartree-Fock 
approximation for the self-energy, equivalent to the Eq. (2.7) rewritten in the particle-particle channel, gives the 
usual BCS result for the Green's function [33]. In the case of DDW order it corresponds instead to the mean- field 
Green's function usually considered in the literature [31, 34-41]. 



A. The electrical conductivity and the conductivity sum rule 



The optical conductivity can be calculated from the electromagnetic response kernel 

2f M „(q, ifl m ) = -'WW 1 _ <W + n Mi/ (q, iO m ), (2.8) 
where Il^^q, iQ m ) is the correlation function 



n /1 „(q > in ro ) = -jf dre^ T (T r ^(q,T)i, y (-q,0)). (2.9) 

Here tu is the diamagnetic (or stress) tensor, r is imaginary time, (3 = 1/T, and f2 m = 2-kjtiT is the bosonic Matsubara 
frequency. The index \x = (i,0) with £ = 1,2 indicates spatial and time components respectively, so that the particle 
current operator j M (q, r) = (ji(q, T),jo(q, r)) consists of the particle current density, Jj(q, r) and the particle density, 
jo (q, t) . As usual, the particle current and the diamagnetic tensor arc defined as the first and second order derivatives 
of the Hamiltonian if (A) in the presence of the vector potential A with respect to A itself [12]: 



H(Ai) « H(0) J2 



e 2 

eA i (T J )j i (T S )-—A$(T S )T ii (T J ) 



(2.10) 



so that the total current density is expressed as Ji(r) = —SH/8Ai(r) = ejj(r) — e 2 Tu(r)Ai(r), and by evaluating 
(Ji(q)) within the linear response theory [12, 33, 42], one obtains J M (<?) = e 2 K fJiU (q)A„(q) with the electromagnetic 
kernel (2.8). Then using that A(w) = E(u;)/i(w + iO), where E is the electric field, one finally arrives at the Kubo 
formula 

. 2 Ku(q = 0,u) . 2 < t« > -njj(q = 0,w) 

cr(cj) = — ie ; — = ze ; : , 2.11) 

V ' V(uj + i0) V(tu + i0) ' V ' 

where the standard analytic continuation iQ m — » uj + iO was made. To avoid confusion, along the paper we will indicate 
the imaginary bosonic frequencies with if2 m , the imaginary fermionic frequencies with iuj n and the real frequencies 
with oj. Since an isotropic system is considered we can omit the index i as done in the LHS of Eq. (2.11) and in what 
follows. 



5 



Taking the real part of (2.11), one obtains Rea(u>) = (ire 2 /V)8(w)[< r > -ReII(0,a;)] + (e 2 /V)Imn(0,w)/w. In 
the presence of disorder the coefficient of the S(u>) vanishes, so that ReII(q = 0, ui — > 0) =< t > and only the regular 
part of a(u>) survives. As a consequence, one usually defines the optical conductivity only through the imaginary part 
of n(q = 0,w): 

„ , . e 2 ImII(q = 0, uj) 

Rea(cj) = ^ (2.12) 

v uj 

so that using the Kramers-Kronig (KK) relations for II(q = 0, uj) one can derive the well-know sum rule: 

f°° , XI e 2 f°° Imn(q = 0,uj) , Tre 2 , x ire 2 , 

W(T) = / Rea(uj)duj = — / — —dui = — R C n(q = 0,cj = 0) = — <t>. (2.13) 

J-oo V J-oo UJ V V 

The form of if (A) itself depends on the microscopic model and thus on the way the vector potential A enters the 
Hamiltonian of the system. When a continuum model is considered instead of Eq. (2.1), the kinetic term is expressed 
as / (— V) 2 /2m and A is inserted using the minimal coupling prescription — iV — ► — iV — eA. For lattice systems the 
equivalent of the minimal coupling prescription is the so-called Peierls ansatz [7, 8, 12], which corresponds to insert 
the gauge field A in eq. (2.1) by means of the substitution Cj — > cie~ le ^ A dr . I n this case, it is clear that when the 
interaction term of the Hamiltonian is a density-density interaction, as in Eq. (2.1), only the kinetic hopping term 
is modified, while the interaction term is gauge invariant (GI). As a result, the current/density operator and the 
diamagnetic tensor can be expressed (for small q) as: 

RBZ 



3l*(*l: t) = Yl l V( k ) C Lq/2/k+q/2a = N jl ^-q/2^( k - q/ 2 > k + q/2)Xk+q/2' q -> ( 2 - 14 ) 

d 2 e k 



k.tr k,cr 



dk 2 

k,cr 1 



n Ka (2.15) 



where 



w M (k) = (t£,l), 7 M (k-q/2,k + q/2) = (^<7 3 ,ao), q^O, (2.16) 

and (v£)i = dsk/dki is the Fermi velocity [43]. Note that if a quadratic band dispersion = k 2 /2m is assumed, 
the tensor Tu reduces to n/m, where n is the total carrier density, so that Eq. (2.13) reduces to the f-sum rule (1.2), 
which is temperature independent. Instead, for a tight-binding nearest neighbors lattice dispersion, according to the 
definition (2.15), tu is proportional to the kinetic energy, and the sum-rule (1.4) is recovered. Observe that formally 
the sum rule (2.13) always requires the integration up to an infinite cut-off energy. Nevertheless, an intrinsic finite 
cut-off energy is provided by the energy scale below which a given model can be considered as a good approximation 
for the real system. As a consequence, while the full f-sum rule is always satisfied at enough large energy scales, the 
restricted optical sum rule relative to a given tight-binding interacting model is expected to hold only below some 
intrinsic energy scale, whose definition is not universal. We would like to stress that the definitions (2.14)-(2.15) 
follow from the Hamiltonian (2.1) once that a gauge-invariant form is chosen for the interaction term. However, 
this assumption is invalid when for example "occupation modulated" hopping terms are present [14], or when an 
"effective" interacting model is considered, in a sense that we will specify below (see Sec. IV). 



B. Gauge invariance and the sum rule 



The derivation of the sum rule presented above is rather formal, and does not allow one to understand that the 
sum rule is just a different way of stating the gauge invariance of the theory. To gain a deeper insight into the relation 
between these two aspects, it is useful to consider here the sum-rule derivation presented in Ref. [42]. The starting 
point is the observation that in a GI theory there is a gauge freedom to choose whether the applied electric field 
E = —dfA — V</3 is included in the Hamiltonian (2.1) either via the vector potential A (cp = 0) or by considering a 
scalar potential <p (A = 0). Obviously, the conductivity derived from two equivalent Hamiltonians H(A) and H(<p) 
must be the same, but this is only guaranteed by the charge conservation 



e5tjo(q, t) + ieq ■ j(q, t) = 0. 



(2.17) 



6 



The proof considered in Ref. [42] that cr(u>) derived from H(A) and H(ip) are the same is based on the identity 

r°° ire 2 1 

/ dwRea{u) = —lim-(\j (c l ,t),j i (-c l ,t)}), (2.18) 

which is obtained by using the charge conservation (2.17). For example, substituting in Eq. (2.18) jo{<\,t) = 
Ek, J c L q /2,/k+q/2.» and thc free-electron expression j (q, t) = (l/«i)Ek, J kc [- q /2, J c k+q/V. corresponding to 
£k = k 2 /2m, returns the full f-sum rule (1.2). 

Another way to state the relation between the sum rule and the GI uses instead the generalized electromagnetic 
kernel (2.8). As discussed in Ref. [33] with reference to the SC case, the requirements of charge conservation 
{q^Jfj-iq) = 0) and invariance of the theory under the gauge transformation A^q) — > A^(q) + iq fl A(q) are fulfilled 
when the condition 

q„K^(q) = K llv {q)q v = 0, q= (q,u) (2.19) 

is satisfied. In particular, thc following relation must hold: 

n ii (q^0,w = 0) = (r ii ). (2.20) 

This equality states that the diamagnetic term is canceled out by the static limit (u> = 0, q — > 0) of the (real) part of 
the current-current bubble, while deriving the Eq. (2.13) we used the relation between the dynamic (q = 0,cj — * 0) 
limit of the bubble and the stress tensor. However, in deriving Eq. (2.13) we assumed the presence of disorder, whose 
role is crucial in restoring the equality between the static and dynamic limits of the current-current correlator II(<j). 
Indeed, while in a clean system thc dynamic limit of thc bubble vanishes, in the presence of disorder it coincides with 
the static limit, which in turn is equal to the diamagnetic term: Ren(tj — > 0, q = 0) = Rcn(cj = 0, q — > 0) =< t >, 
and then Eq. (2.13) follows. 



C. Ward identity and vertex function 

The advantage of the derivations (2.18) and (2.20) of the sum rule is that they show explicitly that it has to be 
regarded as a consequence of the charge conservation. Moreover, it allows one to see that once a given approximation 
is used in evaluating the current-current correlation function, it also fixes the sum rule that will follow from such an 
approximation. However, a quite difficult task is to implement an approximation for both the Green's function and 
the current-current correlator which preserves the condition (2.19), necessary for maintaining thc GI of the theory. 
In particular, when the Hartree-Fock self-energy (2.7) is used and the bubbles n M „ are evaluated in the lowest-order 
approximation: 

njZ? (q, ifim) = Tr[G(k - q/2, iiu n + ifi m ) 7([1 (k - q/2, k + q/2)G(k + q/2, iw n ) 7l/ (k + q/2, k - q/2)], (2.21) 

k 

the GI is not in general preserved, as it is know for SC and as we shall see explicitly in Sec. IV in the case of DDW 
(the factor 2 in the previous equation is due to the spin summation). A general field theoretical approach that 
solves the difficulties with charge conservation and gauge invariance, originally present in the mean-field (bare vertex) 
formulation of the BCS theory, was developed by Nambu [32] and discussed in detail in Chapter 8 of [33], so that 
here we only introduce the main definitions and stress the points necessary for the consideration of the DDW state. 

As shown in [33], thc current-current correlator, defined above in Eq. (2.9), can be expressed in terms of the full 
Green's functions (2.6), the bare vertex 7 M and the full vertex function r„ as follows 

n M „(q,ifi m ) = -2^Tr[G(fc_) 7 ,(k_,k + )G(fc + )r AI (fc+,fc_)], (2.22) 

k 

where k + = (h + ,iLu n + iQ m ,), fc_ = (k_,icj„, ) with k± = k± q/2. The important property of the current-current 
correlation function (2.22) is that the condition (2.19) is preserved whenever the vertex function satisfies the generalized 
Ward identity (GWI): 

g„r„(p +> p_) = G-\ P -) - G- 1 {p+). (2.23) 

The GWI is nothing but the charge conservation law (2.17) rewritten using the Greens' and vertex functions. If the 
Green's function given by Dyson equation (2.6) is evaluated within the Hartree-Fock approximation (2.7), then the 
vertex function satisfying the GWI is also the solution of the following integral equation: 



r>+,P-) - 7m(p+> p-) + E G(fc+)r A1 (fc + , k_)G(k_)v( P - k). 

k 



(2.24) 



7 



The analytical solution of Eq. (2.24) cannot be easily determined, except that in the static limit, when I\ is given by 

r , M . , (p . p) + <m . JJLM _ c-Hvf-^c-^ ,2. 5 , 

Indeed, if one puts q = in Eq. (2.24) (which corresponds, as usual, to the static limit uj = 0,q — > when analytical 
continuation iD. m — > + iO is made), by means of the previous relation one obtains: 

T,(p,p) = n(p,p) + £G(fc)r 4 (M)G(fc)U( P -k) = 7i (p,p) + £fffiy( p -k) 

fc fe 1 

= 7,(p,p) - E G ( fc ) 9 \ k) = 7<(p,p) + ^£ G (*Wp- k ) 

= 7^) + ^- (2.26) 

Here we used the fact that the potential V is non-separable, viz. it depends on the difference p — k, as it is expected 
for a GI density-density interaction. Observe also that this result can be obtained directly from the GWI (2.23) by 
taking the limit oj = 0, q — > 0. For example, one can easily check that WI (2.25) is satisfied for the free electron 
Green's function (2.5) taken together with the bare vertex (2.16). It is worth noting that in the case of SC the behavior 
of the vertex function at zero frequency and momentum is completely different, and indeed Ti(p,p) is divergent as 
the inverse of the phase-mode dispersion [32, 33]. Indeed, the equivalent of Eq. (2.23) contains for the SC case a 
combination of Green's functions and Pauli matrices that cannot be reduced to the derivative of G _1 as in Eq. (2.25). 
Here, however, the equivalent of the gapless phase mode is not present, because there is no Goldstone mode when a 
discrete symmetry is broken, and Ti(p,p) turns out to be finite. 



D. Symmetrized expression for T = dc conductivity 

In practice, since the exact expressions for G and r, are unknown, the consistency of an approximated calculation 
of the conductivity can be guaranteed if the approximated expressions for G and I\ satisfy the GWI (2.23). Observe 
that what we obtained in (2.26) is the limit u> = 0,q — > of T, but in the calculation of the optical conductivity it 
is the opposite limit which is needed. However, at least in the presence of impurities, or at T — 0, the static and 
dynamic limits commute. Unfortunately, a generalization of the result (2.26) to finite frequency cannot be obtained 
from the equation (2.24) for a generic potential, by means of, e.g., a perturbative method. Since our final task is to 
find an approximation for the optical conductivity which allows us also to estimate the corresponding sum rule, let 
us analyze the utility of the result (2.26). First, we note that the knowledge of the vertex function at zero frequency 
allows one to find an exact result for the dc conductivity at T = 0. To show this it is convenient to think of 2 x 2 
matrices A and B as being represented by two column vectors of 2 x 2 matrix elements and rewrite Tr of the matrix 
product as the scalar product: 

Tr[AB] ee ^2(X) aP (B)p a = A-B. (2.27) 

a/3 

Accordingly, by introducing the vector 

(g(k+,k-)?(k+, fc_)) aj8 = £ G aj (k + )Gs l 3(k^)T J s(k + , *_), (2.28) 

7(5 

we can rewrite correlation function (2.22) as follows: 

f d 3 k 

KiM = -2 J ^3e(fc+,fc-)lMfc + ,M ■7;(k-,k + ), (2-29) 

where since we are considering the T = case we have an integration over the real frequency instead of the Matsubara 
sum and the argument of the polarization operator is q — (q, oS). 

The dc conductivity is determined by the imaginary part of the derivative of the correlation function, which in turn 
is given by: 

an li (q = o,w) 



d 3 k 



=° (2.30) 

J [QL (K , *_)r< (k+, k- ) ■ 7, (k_ , k+ ) + g(k + , k-)r' ui (k+ , k- ) • f 3 -(k_ , k+) 



w= 0,q=0 



8 



where 9' U ,T' U indicate the derivative with respect to u>. The expression (2.30) can be further simplified by using the 
equation for vertex (2.24) taken at T = and its derivative with respect to lu: 

%(k t k)=f i (k+,k-)- J -^g(p +1 p-)fi(p + ,p-)VOii-p), q = (2.31a) 

f' ui (k + ,k-) = J ^[^(p +1 p-)f i (p + ,p_) + e?(p +1 p-)f^(p + ,p_)]y(k-p). (2.31b) 
Substituting from Eq. (2.31a) in Eq. (2.29) and using (2.31b) we obtain 



duo 



U1=0 



f d 3 k 

/ 7^3 ^(fc + ,*_;q=0)r i (A !| fc).r i (A !l fc) . (2.32) 



Our derivation is similar to the derivation of the symmetrized expressions for the derivatives of the polarization 
operator considered in Ref. [44], where also the derivative of Bethe-Salpeter kernel enters the analog of Eq. (2.31b). The 
useful property of the representation (2.32) for 11^ is that it contains two full vertex functions IV The corresponding 
expression for the dc conductivity er^c coincides with the result derived by Langer [Eq. (4.8) of Ref. [45]] in early 60s 
using a completely different approach, consisting in introducing a symmetric bubble 

rp RBZ 

Il^ m (in m ) = -2— J2 Tr[G(k, iuo n + ifinOl^ft, fc)G(k, iwnjr^fc, *)], (2.33) 

obtained by using two corrected vertices, evaluated at zero external frequency, and whose derivative at zero frequency 
and temperature coincides with the result (2.32). Then in the limit T — > the leading term of Langer's expression 
for the dc conductivity is obtained from (2.33) via 

e 2 

a dc = Urn Reer(w) - -Im d^ m (in m -> u;)]| . (2.34) 

to — *0 y 

In Eq. (2.26) the vertex function with coinciding fermion momenta and energies, k + = fc_ = k is related to the self- 
energy E(fc) by the WI (2.25) [39, 45]. Thus one can immediately see that whenever E(fc) depends on the momentum 
k, the dc conductivity (2.34) would be different from the value obtained using the bare bubble (2.21). 

From the previous considerations one can argue that, in the absence of a solution for the vertex function T at finite 
frequency, a better approximation for the conductivity in the DDW state is provided by the bubble (2.33), which 
gives at least an exact result for the dc conductivity at T = (see also Ref. [39]). In other words, by evaluating 
the symmetric bubbles (2.33) at finite frequency one can still capture the behavior of o~(u>) at small lu. At the same 
time, we do also expect that this assumption will lead to a new result for the sum rule (2.13), because the symmetric 
bubble (2.33) is no more connected to the diamagnetic term (2.15) by any relation. However, as we shall see in the 
next Section, the sum rule for the bubbles (2.33) can be obtained analytically by using the analogies between the 
results discussed up to now and the properties of the reduced Gaussian model, where the vertex equation admits the 
solution Eq. (2.26) at all frequencies. 

III. VIOLATION OF THE GI WITH THE BARE VERTEX IN THE DDW STATE 

A. The mean-field DDW Hamiltonian 

The previous discussion was generically referred to any system displaying a particle-hole instability at the wave 
vector Q. However, in the Hartree-Fock approach one usually selects a particular form for the mean-field Green 
function G and then solves the self-consistency equation corresponding to implement the Dyson equation (2.7). In 
the DDW case, one approximates the general interacting Hamiltonian (2.1) with the model: 

Hi = u; d( k ) u; d( k/ ) c L c k+Q<74'+Q CT ' c k'<T', (3.1) 

k.k' 

where Wd(k) = (cosk x a — cosk y a)/2. By defining iD a = —(Vo/N) ^2 ka uv(k) < c^ + Q a Ci c . a > wc obtain the following 
mean-field DDW Hamiltonian 

RBZ 

H = X^[( £ k - M)cL c k<r + i-DkC^Ck+Qa] = xL [ £ k03 - D k a 2 - /x] Xk<r, (3.2) 

k,<r k,cr 



9 



where Z?k = DoWd(k) is the gap, known as the DDW gap [27], arising from the formation of the state with circulating 
currents below a characteristic temperature Tjjdw [46]. This Hamiltonian corresponds to an explicit solution of 
Eqs. (2.6)-(2.7) with 

E(k) = -D k a 2 , (3.3) 

so that the full Green's function (2.6) reads: 

G _1 (k, iw n ) = (iui n + n)a - e k a 3 + D^a 2 . (3.4) 
The corresponding self-consistency equations for the order parameter Dq and for the chemical potential fi read: 



N 2^ 



RBZ 2 ( 

A " J [/(£-,k) - /(£+*)] = 1, (3.5) 



k k 

2 RBZ 



„ £[/(£- 1 k) + /(£f,k)]=n ) (3.6) 



k 



where i?k = \/^k + ^k> anc ^ £±, k = — ^ are * ne * w0 excitation branches associated with the formation of DDW 
order, which breaks translation symmetry. Observe that to obtain the Eq. (3.6) we used the fact that the occupation 
number n^ a in the DDW is given, according to the Green's function (3.4), by rtker = (l/2£'k)[-E'k(/(£+) + /(£-)) + 
£k (/(£+) — /(£-))]■ This allows us also to evaluate the diamagnetic term (2.15) and the corresponding sum rule as: 

(35m =< * >= 4 R fi^ - m 

where we used the fact that (d 2 e-^/ dkf) = +2ta 2 cos(kia) and we omitted the explicit dependence of e,E,£± on k. 
At this level we have performed an approximation on both the self-energy and the Green's function of the original, 
interacting system. To obtain now a GI approximation for the optical conductivity, i.e. an approximation which 
gives Eq. (3.7) as the integral of cr(uS), we should also evaluate the vertex function (2.24). Indeed, as we show 
with an explicit calculation in the next section, the bubble (2.21) with a bare vertex 7 violates this requirement. In 
general, if the optical conductivity cannot be calculated by means of the exact vertex function (2.24), but a different 
approximation is used, one cannot expect any more to find Eq. (3.7) as the corresponding sum rule, but this has to 
be calculated explicitly, as we do in Sec. IV. 

Before showing the details of this calculation we would like to comment on the relation between the microscopic 
interaction (2.1) and the approximated one given in Eq. (3.1). If one restrict in the interacting part of Eq. (2.1) the 
sum over nearest-neighbors sites one can easily show that Hi nt can be rewritten as: 

H int = -V ^ ^a(k)u'a(k / )Ck -C3k+qff' c k'+q < r' c k'o-! ( 3 - 8 ) 

where the factor (1/4) ^2 S e*' k " k ) coming from the sum over nearest-neighboring sites 5 has been decomposed in 
the two-dimensional basis of wave functions w a (k), which includes the sum of contributions from several channels 
displaying different symmetries with respect to the discrete rotation group for the square lattice. One can see that 
even selecting only the d-wave channel and the contribution at q = Q Eq. (3.8) does not contain only the coupling 
in the particle-hole channel of Eq. (3.1), because the spin structure in Eq. (3.8) and Eq. (3.1) are different. This 
problem does not exist if the original microscopic model is given by a current- current interaction, as the formation of 
a DDW state would naturally require: 

Hint = — ^2 cl^Cicr'Cj^^jcr = — c icr C ja c ]cr' C i"' ■ (3-9) 

<i,3> <ij> 

Observe that: (i) Eq. (3.9) is still gauge invariant, since the Peierls transformation does not depend on the spin index; 
(ii) the self-energy for the interaction (3.9) is still given by Eq. (2.7), with V(q) = 2Vo. In the following we will never 
face with the problem of solving explicitly Eq. (3.5) for a given microscopic interaction. However, it is worth noting 
that Eq. (3.1) can be directly derived by selecting a specific channel of a microscopic GI model. Other examples can 
be also found in Rcf. [20-25]. 



10 



B. The current-current correlation function evaluated with the bare vertex 7 



To evaluate the bubbles (2.21) it is useful to introduce the spectral representation for the Green's function (3.4): 

G(k,iu} n )= [ dz Ai ^' ^ (3.10) 

with the spectral function 

A(k, z) = — S(z - [i- Ev) + — S(z -fi + Eu). (3.11) 

The correlation functions (2.21) can then be written as 

nW(q,ifi m ) = ~Y, I dzidz 2 Tr[A(k + , ^) 7 ^(k + , k_)A(k_, z 2 ) 7v (k_, k + )] {^'[^ (3.12) 

k 

which gives, according to (3.11), the following current-current correlation function: 



RBZ 

n^(q^ m) = -l£«) 2 

k 

+«) 2 

(3.13) 

where D± = L> k±q / 2 , £± = £ k±q / 2 , E± = J B k ± q /2- 

The issue then arises of the relation between the approximation (3.12) for the correlation function and the sum rule 
(3.7). Let us consider again the GI relation (2.20). When the static limit of the current-current bubble is considered 
(corresponding in Matsubara formalism to put q = in the expression (3.13)), we find that I[( 7 )(0,0) is real and 
given by 

_ RBZ i F\2j-)2 1 F\2 2 

nl 7) (o,o) = ~ £ ^r-\f{S+) /(£-)] + ^^[fiM + /'(£-)]• (3.u) 

k 

The usual procedure used to demonstrate that 11(0,0) cancels out < r > given by Eq. (3.7) consists in integrating by 
parts the term in 11(0, 0) which contains the derivative of the Fermi distribution /'(£±) [12]. In order to do that one 
would need a term like df(£±)/dki — ±f'(£±)(ev£. — Dv£)/E, which in addition to the Fermi velocity, v£, contains 
also the DDW gap velocity, v£. = —dD^/dhi. However, as one can easily see, the second term of Eq. (3.14) does not 
contain any contribution proportional to , so that the gauge-invariant relation II(q — > 0, oj = 0) =< r > cannot be 
satisfied with the bubble n*- 7 -'. 

According to the discussion of the previous section, the GI can only be restored when the vertex corrections are 
included in the correlation functions. Observe that the static-limit result (2.26) reads in the case of DDW state: 

Ti(k, k) = 7i (k, k) + V&2 = <<7 3 + <o-a = ViQt), (3.15) 

where V^(k) is the generalized velocity representing the q = limit of the vertex function. It is important to stress 
that in deriving Eq. (2.26) for T^fc, k) it was crucial to keep the translation invariant form V(p — k) of the potential 
until the end. This point has been often overlooked in the literature, at least while discussing the corresponding 
problem for the superconducting case [33]. If one used in Eq. (2.24) defining the vertex function the approximated 
form V(k — p) sw VoWd(k.)wd(p) , which is appropriate for selecting only the d-wave channel in the self-energy (2.7), 
the result (3.15) could not be obtained. Indeed, a single-channel separable potential makes the interaction term of the 
Hamiltonian not gauge invariant, and then it would contribute to both the current operator and diamagnetic term, as 
we will discuss in the next Section within the context of the reduced model. However, once that the result (2.26) has 
been established, and all the intermediate steps have been performed in respecting GI requirements, we can definitely 
select from the self-energy (2.7) only the <i-wave channel. As a consequence, if S(p) is approximated as in Eq. (3.3), 
the result (3.15) follows. 



f(-V + E + )-f(-LL + E_) , /(— /x — E + ) — /(— /i — E_) 



E + - E- - ifl m E+ - E-+ i£l m 

/(- M + E+) - /(— A» - E-) /(-/x - E+) - /(-/x + 



E_ - in r 



E+ + £L + 



1 - 



£+£_ - D+DJ 

E+E- _ 

£+£_ - D+DJ 

E+E- 



11 



The Eq. (3.15) can also be obtained from the generalized WI (2.23). Indeed, at small q the difference G l (p) — 
G _1 (p + q), where G is defined in Eq. (3.4), is given by 

q^„ = G-\k) - G-\k + q) = -in m <j + v£ ■ q^3 + vg a% (3.16) 

Since, according to the definitions (2.16), the bare vertex satisfies qpTIp, = —i£lm&o + w k ' f° r = we can 
find again that the static vertex (3.15) satisfies the WI (2.23). Note that from the WI (3.16) one can be tempted 
to generalize the result (3.15) for all fi TO ,q ~ 0: however, one cannot exclude that an additional term with zero 
space-time divergence can be added to the solution (3.15), still satisfying Eq. (3.16). 

According to the discussion of Sec. II-D, one can try to use the result (3.15) by evaluating the optical conductivity 
with the symmetric bubble (2.33). In the specific case of the DDW order, this would correspond to evaluating the 
following current-current correlation functions: 

„ RBZ 

ng w (q,iO m ) = -2- J2 Tr[G(k_,io;„ + if2 m )y i (k)G(k +J ia;„)^(k)]. (3.17) 

As it was explained in Sec. II D the ansatz (3.17) guarantees the correctness of the dc conductivity, and in general 
can be used to study the low-frequency conductivity. Nevertheless, one can check that the bubble n? w (0) is not 
compatible with the diamagnctic tensor (3.7), violating again the GI condition (2.20) checked above for u\]\ The 
origin of this violation is obvious, viz. instead of the asymmetric bubble (2.22) with one full and one bare vertex 
that would maintain the GI condition (2.20), we used the symmetric correlation function (2.33). Thus the issue arises 
whether the diamagnctic tensor (r) can also be modified to become compatible with the bubble (3.17). As we shall 
see in the next section, the diamagnetic tensor and sum rule corresponding to the approximate bubble (3.17) can be 
obtained without further assumptions by analyzing the properties of the reduced Hamiltonian (3.2). 



IV. THE REDUCED MODEL 



An approach often proposed in the literature to deal with the DDW state is that to consider directly the mean-field 
Hamiltonian (3.2) as the starting point [31, 34-40]. The idea is that at low energy the reduced model (3.2) captures 
the important physics of the system, so that one can consider it as a starting microscopic Hamiltonian, describing 
non-interacting quasiparticles. In this case the Green's function (3.4) does not provide any more an approximation, 
but it is the correct one for the solvable, quadratic model (3.2). Since this Hamiltonian describes non- interacting 
quasiparticles, it can be solved exactly and the corresponding conductivity is given by the bare bubble. This point of 
view was taken in [31] where an unusual form of the optical-conductivity sum rule was obtained. One can notice that 
any distinction in the Hamiltonian (3.2) in the total energy between a kinetic and a potential part, as can be done 
for the Hamiltonian (2.1), becomes somehow ambiguous, so that the result of [31] is not surprising. 

In what follows we compare this picture with the traditional one, and show that since the dc conductivity calculated 
in both approaches appears to be the same, one can also estimate the low-energy sum rule of the microscopic model 
(2.1) by considering the one realized in the reduced model (3.2). 



A. The diamagnetic tensor, current operator and the sum rule for the reduced model 

Let us now consider the Hamiltonian (3.2) as the starting microscopic model and analyze how all the considerations 
made in Sec. II can be applied in this case. Since the Hamiltonian (3.2) describes non-interacting quasiparticles, it 
is straightforward to calculate the current-current correlation function and the electrical conductivity, because in the 
absence of an interaction term the Eq. (2.24) for the vertex has a trivial solution r M (p + ,p_) = 7 M (p + ,p_), where 
^(p+jP-) is the bare vertex for the model DDW Hamiltonian (3.2). Nevertheless, one should be careful and take into 
account that this vertex is different from the bare vertex (2.16) for the Hamiltonian (2.1). This can be understood by 
deriving the particle current operator compatible with the conservation law (2.17) and with the equations of motion 
for the operators c and c' , [36, 37, 39, 40] 

j(q, *) = K C U/2" Ck +l/ 2 » - iJ, k4-q/2/k + Q+q/2 I r] , (4.1) 



The first term of the previous expression relates as usual the particle current to the band velocity v£ . The second 
term, which only appears for non- vanishing Dq, takes into account the contribution of the orbital currents to the 



12 



electrical conductivity, arising when the DDW order is established. Observe that in the reduced model (3.2) the term 
proportional to Do appears as an additional, temperature dependent band, which couples k and k + Q electrons, 
and as a consequence a corresponding term appears in the definition of the current. By rewriting the electric current 
operator (4.1) using the spinors (2.2), one has 

RBZ 

j*(°>*) = ]vExLK(k)xk, (4-2) 

and, accordingly, the bare vertex reads [39]: 

7 Al (k-q/2,k + q/2) = (^(k),a ), q 0, (4.3) 

where Vi(k) is the generalized velocity defined in Eq. (3.15). Substituting the bare vertex (4.3) and the Green's 
function (3.4) in the Ward identity (2.25) one can easily see that it is satisfied. Moreover, since for non-interacting 
quasiparticles the full and bare vertex functions coincide, the correlation function of Eq. (2.21), evaluated with 
the bare vertex 7 of Eq. (4.3), has two properties: (i) it is the exact one for the quadratic model (3.2); (ii) it coincides 
with Ilfj DW in Eq. (3.17), which is an approximation for the full model (2.1). As a consequence, the sum rule 
corresponding to the bubble Ilfj DW can be obtained by the knowledge of the stress tensor for the reduced system. 
Observe that the current operator (4.1) is also obtained when the Peierls substitution is performed directly in the 
reduced model (3.2). As we discussed in Sec. II A, after the Peierls substitution both the current operator and the 
diamagnetic tensor can be derived from H(A), according to Eq. (2.10). As a consequence, in the reduced model not 
only the current operator but also the diamagnetic tensor tu is modified, containing an extra term for Dq ^= [31], 



~2N [ £ k(cL C k.) + «Ac(4 CT Ck+Qa>] • (4.4) 
kcr 



When the operator averages are evaluated, or analogously Eq. (2.18) is used, one finds that the sum rule for the 
reduced model is: 



(neW/V) ~~N^ E *Vti+*> fti-*)l> 



(4.5) 



k 



where and £±,k were already defined after Eq. (3.6). Eq. (4.5) was derived using the fact that d X:V v£ = 2ta 2 cos k x ,yd 
(and d a , v vg = ±(A>/2)a 2 cos k x ^yO) , and it reduces to Eq. (1.3) for Dq — 0. 

Once more, the result (4.4) is consistent with the GI for the reduced model. Indeed, if the bubbles U DDW (q, iQ m ) 
(3.17) are evaluated in the static limit, instead of the result (3.14) for n^ 7 ) one has: 

nr(o,o) = -iff {vfD +f £) \ m) m-)} + ( " fg ~f J) V (£+) + n-)] (^) 

k 

If now one integrates by parts the second term of Eq. (4.6) one finds that: 

4 E « + ) + /-«-)] - £ £[/«+) - . 

k k 



2 ^ e) m+) m _ )} _ 1 ^ E[m+) m _ )}> 



and as a consequence the GI relation H® DW (0,0) = (tu) with (tu) given by Eq. (4.4) is satisfied, as expected when 
the exact vertex T in included in the bubble. 

A comment is in order now about a third possible approach proposed in the literature [38] for the analysis of the 
reduced model (3.2). By rewriting the quadratic Hamiltonian (3.2) as H = ^2 kcr \^ a H^x^ a the matrix can be 
diagonalized by means of an unitary transformation [7k, H = U^A^U^ where Ak = diag(£+,£-). According to our 
definition (4.1), the current is derived from H^(A), so that it corresponds to Jddw = (V^O Ylka Z Xwai^^)^^ = 
(l/iV)^ k(T Xko-^k(t^kAkC/j^)xkcr (see also [40]). Let us introduce the spinors ^ a — U^Xka which diagonalize the 



13 



Hamiltonian matrix H, H = Y^ka Z V'ko-Ak'i/'kCT- Then, by making the assumption that the gauge field couples by 
Peierls ansatz not to Xko- but to the new quasiparticle operators V'kc, one would calculate the current starting from 
Ak(A), so that the current, the diamagnetic term and the static limit of the current-current bubble would be defined 

as [381: 



RBZ 



RBZ 



3QP = jr V'k CT c' k A k V'k ( T = ^ xi a U k {d k A)U^x^, 

kcr 

^ RBI 

< TU,QP >= T7 £ 



RBZ 



n.„, Q p(o,o) = -l 

k<7 



9|+ 
dki 



/'(€+) 



dkl 

2 



/'(£-) 



Observe that this approximation is still GI in the sense that it is easy to see that tu^qp and Ylu t Qp(0) defined above 
satisfy the condition (2.20). However, this approximation has no relation with the microscopic starting model, in the 
way we explained in Sec. II. For this reason, we do not comment further on this approach, and we analyze instead the 
result obtained with the current operator (4.1) and the bubbles (3.17), whose correspondence with the microscopic 
model we established above. 



B. Temperature dependence of the spectral weight 

Once that we clarified the different approximations used in deriving the two sum rules (3.7) and (4.5), let us discuss 
the outcomes of these two approaches as far as the temperature dependence of the spectral weight is concerned. 
Two observations should be kept in mind: (i) the overall variations of the spectral weight in the DDW state are not 
expected in general to be large if quite small gap values D(0)/t <C 1 arc considered; (ii) in Eqs. (3.7) and (4.5) the 
temperature variation of both the gap and the chemical potential fi contribute to the shape of W(T). In the case 
of free electrons, the variation of n(T) is almost negligible compared to the temperature variation of the occupation 
number, given by the Fermi function. Indeed, even considering the temperature variation of n{T) the result (1.4) 
is only modified by terms of order X" 4 . In the case of Eqs. (3.7) and (4.5) also the band structure is varying in 
temperature, and it is important to keep track of this by solving at each temperature the self-consistent equation 
for the chemical potential. Here, instead of solving explicitly Eq. (3.5), we adopt a general mean- field temperature 
dependence for Dq(T) = D(0)g(T/Tp>Dw)} with g(x) = (1 — i 4 /3)vl-a?, as shown in the left panel of Fig. 1. In 
the right panel of Fig. 1 we also present the temperature dependence of the chemical potential in the DDW state: as 
one can see, below Tp,jyw there is an inversion of tendency of fi(T) due to the opening of the gap. The temperature 
dependence of the spectral weight in the DDW state according to Eq. (3.7) (W(D,T)) and (4.5) (W DDW (D, T)) is 
reported in Fig. 2, where also the tight-binding spectral weight (1.4) (W(T)) is shown for comparison. Here we used 
a small gap value, D(0) — 2.hToDW — 0.3f, and doping 8 — 0.1. The influence of the chemical-potential variation 
are evident comparing the right panel of Fig. 2, where W(D,T) and W DDW (D, T) are evaluated keeping fi constant 
(at the value it has in the normal state), and the left panel, where the density is constant. In addition, we see that 
for this value of -D(O) the overall spectral-weight variations are small in the DDW state. However, it is found that 
the definition (3.7) leads to a smooth decrease of the spectral weight below Tddw, in analogy with the results for a 
SC transition, while the definition (4.5) gives an increase. Such variations are quantitatively (but not qualitatively) 
modified if the temperature variations of the chemical potential are not properly taken into account, see right panel 
of Fig. 2. Observe that the relative variations of W(T) between T = 0.16i and T = are never larger than ~ 1.2%, 
and cannot be appreciated on the scale of the figure reported in Ref. [40]. 

Even though a detailed description of cuprates is not the main aim of our paper, we find nevertheless useful to 
compare our results for a choice of parameters appropriated for HTSC. Since on this respect different attitudes are 
present in the literature, we briefly recall here the phase diagram analyzed in Ref. [26] within the more general attitude 
of investigating the consequences of describing the pseudogap state with a fc-space modulated charge density wave. 
In Ref. [26] it was shown that one outcome of this description is the possibility to interpret the leading-edge shift 
observed in photoemission experiments as due to a particle-hole gap. In particular, for a band dispersion with a 
next-nearest neighbors hopping term t' — the holc-pockcts Fermi surface formed by doping the DDW system with 
respect to half-filling is a simplification intended to reproduce the arcs of Fermi surface observed experimentally. A 
simple calculation shows that in such a case the gap measured by ARPES at the M points corresponds approximately 
to Dq — l/^l- As a consequence, Dq(0) and Tddw do not correspond directly to the maximum gap value and the T* 



14 



D (T) (i(T) self-consistent 




0.04 0.08 0.12 0.16 0.04 0.08 0.12 0.16 

T/t T/t 

FIG. 1: (Color online) Left panel: Temperature dependence of the DDW gap according to the function g(x) defined in the text. 
Right panel: (J,(T) in the normal state (dashed line) and in the DDW state (solid line), obtained solving the self-consistency 
equation (3.6) for the particle number (with Do = for the normal state). 

temperature measured by ARPES, but both are quite larger, as shown in Ref. [26] where the values of the DDW gap 
and of the temperature Tddw were chosen to properly reproduce the phase diagram of B42212 compounds. 



H(T) self-consistent n(T) not self-consistent 




0.04 0.08 0.12 0.16 0.04 0.08 0.12 0.16 

T/t T/t 



FIG. 2: (Color online) Spectral weight W DDW (D,T), W(D,T), and W(T), according to Eqs. (4.5), (3.7), and (1.4), respec- 
tively, in units of e 2 iva 2 /V. We used here D(0) = 2.5Tddw = 0.3i and 8 — 0.1. Left panel: results obtained using the chemical 
potential obtained solving the self-consistency equation (3.6) for the particle number with Do 7^ (for W DDW (D, T) and 
W(D,T)) and D = (for W(T)) respectively, see Fig. 1. Right panel: evaluation of W DDW {D,T) and W{D,T) using the 
chemical potential of the normal state. 

In agreement with Ref. [31] we keep here this attitude and use a doping and temperature dependent DDW gap 
D (T,6) = cT DDW (5)g(T/T DDW ), where T DDW {5) = 0.16£[1 - (S/5 ) 4 } vanishes at the critical doping 5 = 0.2 for 
DDW formation, and c = 7 is a fitting parameter [26] . Since the resulting temperature dependence of the sum rule 
(4.5) was already shown in Ref. [31], we just report here for comparison the behavior of the two sum rules W(D, T) 
in Eq. (3.7) and W DDW (D, T) in Eq. (4.5) for this choice of parameters at 8 = 0.16. As it can be seen in Fig. 3, the 
relative variation of the spectral weight below Tjjdw is made now more pronounced, enhancing the differences between 
the two possible approaches followed in deriving the sum rule. It is then clear that the standard sum-rule derivation 
leading to W(D,T) in Eq. (3.7) cannot be consistent with the experiments, since no decrease of the spectral weight 
has been observed in the pseudogap phase of cuprates. The result W DDW (D,T) in Eq. (4.5) is instead resembling 



15 



more closely the experimental findings, in particular if we consider that at this doping level the room temperature 
below which the data in Ref. [1, 2, 4] are reported corresponds to T/t ~ 0.1, so that the overall measured temperature 
dependence of W(T) would correspond in our picture to the DDW result (4.5). Indeed, as we show in the right panel 
of Fig. 3, the W DDW (D, T) evaluated according to Eq. (4.5) still displays a T 2 temperature dependence, but with 
a larger slope, as observed experimentally. This approach would allow one to understand why the spectral-weight 
increase looks like a "standard" free tight-binding model, but with a much larger slope. However, as we shall see in 
the next subsection, the comparison with the experiments is made much more involved when the optical conductivity 
corresponding to the sum rule (4.5) is evaluated. Finally, one can in principle extend this analysis to the case where 
also SC is added, but since also the experimental situation is not clear on this respect we refer to Ref. [31] for a 
discussion about the SC state. 

0.825 

0.82 
0.815 

0.81 
0.805 
g 0.8 
0.795 

0.79 
0.785 

0.78 

0.775 „.,„ 

0.04 0.08 0.12 0.16 0.01 0.02 

T/t (T/t) 2 

FIG. 3: (Color online) Left panel: Spectral weight W DDW (D,T), W(T), and W(D,T) in units of e 2 Tia 2 /V for a choice of 
parameter values appropriate for cuprates (see the discussion in the text). Here we show the results for S = 0.16, corresponding 
approximately to optimal doping, and we calculate the chemical potential self-consistently. Right panel: spectral weight plotted 
as a function of (T/t) 2 : one can see that a T 2 temperature dependence is still recovered for W DDW (D,T) in a wide range of 
temperatures. 




C. The role of a next- nearest neighbors hopping term 

Up to know we did not consider the possibility of a next-neighbors hopping term il in the bare band dispersion 
£k- Indeed, from one side we wanted to simplify the notation while discussing the issue of the relation between gauge 
invariance and sum rule, and from the other side we believe that even when comparing with cuprates the case t' = 
is enough to reproduce phenomcnologically the arc of Fermi surface observed in the pscudogap phase (see discussion 
above). However, for the sake of completeness, we report here briefly the modifications induced in the sum rule when 
a t' term is included in the band dispersion, so that 

£k = Sk +Pk, 
St = — 2i(cos k x a + cos k y a) 

Pk = 4i' cos k x a cos k y a. (4-7) 

In the DDW state the perfect nesting condition is lost due to the t' term, so that £k + £k+Q = 2pk, £k — £k+Q = 2sk 
and the two quasiparticle branches in the DDW state become £±,k = pv — /Lt ± -Ek, where = \/ s k + ^k- As a 
consequence, given the relation (2.13) between the sum rule and the diamagnctic tensor, and the definitions (2.15) 
and (4.4) of the diamagnetic tensor for the original and the reduced model respectively, it is easy to see that Eqs. 



16 



(1.4), (3.7) and (4.5) get modified as: 

W{T) 
(7re 2 a 2 /V) 

W(D,T) 
(ire 2 a 2 /V) 

W DDW {D,T) 
{ne 2 a 2 /V) 



RBZ , y 

s 



k 

B.BZ 



(4.8) 
(4.9) 
(4.10) 



where the explicit dependence on k is omitted. In Fig. 4 we compare again the temperature dependence of the spectral 
weight in the different cases, for t' = 0.3t, S = 0.1, Tjjdw = 0.12t and D(0) = 4Tdehy. Even though the introduction 
of the t' term modifies the temperature dependence of the chemical potential in the normal and DDW state (due to 
the shift of the Van Hove singularity which is now below the Fermi level at 8 = 0.1), the general trend of Figs. 2-3 
is confirmed. Indeed, W DDW (D,T) is larger than W(T) below Tddw-, while W(D,T) is smaller. In particular, it is 
worth noting that apart from possible quantitative differences with respect to the case t' = 0, the exact form of the 
band dispersion is irrelevant as far as the main issue discussed in the previous sections, i.e. the fact that different 
approximations for the current-current correlation functions lead to different results for the optical-conductivity sum 
rule. For this reason, we do not discuss further in the following the role of a t' term, and remind the reader for 
example to Refs. [40, 41], where this issue is investigated in more details. 



t=1,t'=0.3t, 8=0.1, T n 



0.81 
0.805 
0.8 
0.795 

0.79 
0.785 

0.78 
0.775 

0.77 
0.765 





_ ' W Dbw (D,T) ' 




W(T) 




\. W(D,T) 




/ 

/ 











0.02 0.04 0.06 0.08 
T/t 



FIG. 4: (Color online) Spectral weight in the presence of a t' term in the band dispersion. Here we show W DDW {D, T), 
W(D,T), and W(T), according to Eqs. (4.10), (4.9), and (4.8), respectively, in units of e 2 Tva 2 /V. We used here t' = 0.3t, 
Tddw = 0.12t, -D(O) = 4Tddw and 5 = 0.1. The chemical potential is evaluated self-consistently at each temperature by 
solving Eq. (3.6) in the presence of a t' term in the band dispersion. Observe that near Tddw a small decrease of W DDW with 
respect to W(T) is observed, due to the change of chemical potential near Tddw- 



D. The optical conductivity of the reduced model 

As we discussed in the previous sections, one would expect that our result (4.5) for W DDW (D,T) is only valid at 
low energy scales, possibly below the plasma frequency, which can be thought as a general cut-off for any tight-binding 
based description of the system. In principle one could also expect that the low-energy theory (3.2) is valid at even 
lower energy scales, but since at the plasma energy one still finds experimentally strong variation with respect to the 
naive estimate (1.4), it is plausible that a quite larger cut-off holds here for the tight-binding model itself. To analyze 
the dependence of the result (4.5) on the cut-off frequency we need an explicit calculation of the optical conductivity 
obtained with the bubble (3.17). 

By using the spectral representation of the Green's functions the current-current correlation function Iif i DW (3.17) 
can be evaluated in analogy with IL^ in Eq. (3.12), with the bare vertices 7,; substituted by the full one r 2 ;(fc,A:) 
of Eq. (3.15). To take into account the effect of disorder we make the simplest ansatz of substituting the delta 



17 



functions associated to a quasiparticle pole in the spectral representation (3.11) with a Lorentzian of finite width w 
(w = l/(2r tr ), where r tr is the transport time): 



5(z)^M(z) = -^— 5 . (4.11) 



(ew c +Dv F ) 2 



1 it) 

7T Z 2 + W 

As a consequence, after analytical continuation in Eq. (3.12), we obtain: 
a DDW (oj) = 
2ne 2 f jJ{z + uj) - f(z) \ (ev F - Dv D f 

■[M(z + uj- Z+)M{z - i_) + M(z + u- £_)M(z - £+)] 

where v F and v D refers to the component in a given x or y direction. 

As already observed in Ref. [37], and more recently in Refs. [40, 41], the optical conductivity is composed of two 
contributions, due to the splitting of the original single band £k in two new bands £± after the gap opening. In Eq. 
(4.12) the first line describes intra-band excitations (corresponding to the product of two M functions evaluated at 
the same quasiparticle branch), while the second line takes into account inter-band processes. It is easy to see that 
this second contribution is only possible when oj > 2|/i| (at low temperatures). Indeed, when the system is doped 
with respect to half- filling (\/j,\ ^ 0) the smallest energy difference between occupied and unoccupied states in different 
branches is equal to 2\/j,\, and it is realized at the points (±7r/2, ±7r/2) where the energy E^ vanishes and the two 
bands merge. The first contribution has instead a Drude-like shape, as it is shown in Fig. 5, where we report the 
optical conductivity at T = for a system with and without DDW gap. Here we used the set of parameters discussed 
above for cuprates, at S = 0.13. When compared with the free-electron conductivity at the same temperature, one can 
see that the Drude peak is smaller in the DDW state, because part of the spectral weight has been transferred to the 
inter-band processes. To quantify this transfer of spectral weight we integrate numerically the optical conductivity 
a DDW (uj) (4.12), and its analogous u(oj) at Dq = 0, evaluating for a given cut-off frequency oj the quantity: 

n (ddw) {u)) = 2 f U a (DDW)^ dL y } (4 . 13 ) 



JO 

which verifies N DDW (oj oo) = W DDW (D, T = 0) and N(oj -> oo) -> W(T = 0) for the DDW and normal state 
respectively. 

In the inset of Fig. 5 we show N DDW (oj) and N(oj) at T = corresponding to the calculated optical conductivities. 
As we can see, at low energy the formation of a DDW state leads to an overall decrease of spectral weight, since 
intra-band processes are partly suppressed. However, at higher energy inter-band excitations are allowed and the 
spectral weight lost in the Drude peak is over-compensated, giving rise to an overall increase of W DDW (D, T = 0) in 
the DDW state compared to W (T = 0) in the normal state. For a value of t ~ 0.25 eV one sees that in the case of Fig. 
5 the crossing of N DDW (oj) with respect to N(u>) is already satisfied at cut-off frequencies smaller than the plasma 
frequency (w 4t), even though N DDW saturates at higher frequencies. Of course it is evident that the determination 
of the exact frequency at which the sum rule W(T) or W DDW (T) are exhausted depends on the choice of parameters. 
For example, at smaller doping or smaller Do (which both lead to a smaller value of \fi\ in the DDW state, see Fig. 
1) the intra-band processes occur at lower energy, so that N DDW (uj) > N(u>) will be satisfied at lower cut-off energy 
lu. In Fig. 5 we report one of the cases when the cut-off frequency is larger, because the chemical potential shifts 
from /j, = — 0.28i in the normal state to /i = — 0.69£ in the DDW state, due to the large value of the DDW parameter 
(Dq = 0.92i), pushing inter-band processes at relatively high energies. It is interesting to observe that the optical 
conductivity in the DDW state reported in Fig. 5, which was evaluated with the ansatz (3.17) for the current-current 
correlation function, has the same qualitative behavior of the optical conductivity reported in Ref. [41], where the bare 
bubble approximation (2.21) was considered for the correlation function. Indeed, in the bare bubble approximation 
the optical conductivity has the same structure of Eq. (4.12) (i.e. a Drude term plus inter-band processes), but with 
Vd = 0. However, the two approaches lead to two quantitatively different temperature dependences of the sum rule. 
Indeed in the absence of the vd term in Eq. (4.12), coming from the vertex corrections, the spectral weight lost in 
the Drude term when the DDW state is fromed would not be compensated any more by the inter-band processes, so 
that the total spectral weight N DDW (oj — > oo) in the DDW state would be always lower than the spectral weight 
N(uj — > oo) in the normal state, as observed in Ref. [41]. 

In Fig. 6 we report the optical conductivity in the DDW state at several temperature between T = and Todw = 
0.13£ at the doping 5 = 0.13. As one can see, when the temperature increases the inter-band processes shift to 



18 



t=0.25 eV, 5=0.13, T=0 
to (cm* 1 ) 

800 1600 2400 3200 4000 4800 5600 6400 7200 8000 

1.4 



1.2 
1 

3 



0.4 
0.2 



0.5 1 1.5 2 2.5 3 3.5 4 

co/t 

FIG. 5: (Color online) Optical conductivity in units of e 2 na 2 /V at zero temperature for a free tight-binding system and for 
the DDW state at 5 = 0.13 (/i = — 0.694 at T = in the DDW state, w — O.lt were used). For convenience we also report 
the frequencies in cm -1 , as it is customary in the experiments. Inset: frequency variation of N(to) according to Eq. (4.13). 
Observe that at low cut-off energy the spectral weight in the DDW state is smaller than in the normal state. 

t=0.25 eV, 5=0.13 
co (cm" 1 ) 

800 1600 2400 3200 4000 4800 5600 6400 7200 8000 

0.9 
0.8 
0.7 



0.6 



< 0.5 
3 

e 0.4 



0.3 
0.2 
0.1 





0.5 1 1.5 2 2.5 3 

co/t 

FIG. 6: (Color online) Optical conductivity in units of e 2 Tra 2 /V at various temperature for the DDW state at 8 = 0.13, 
Tddw = 0.13t. In the upper x-axis the frequencies are reported in cm -1 . Inset: comparison between the temperature 
dependence of the integrated spectral weight N DDW (lom) and N(lom) for the DDW and normal state respectively. The cut-off 
is lom = 6t. 

lower frequency, due to the decrease of the absolute value of the chemical potential (see right panel of Fig. 1). As 
a consequence, the spectral weight is transferred again towards the Drude peak, and the overall balance of spectral 
weight leads to a decrease of W DDW (D, T). In the inset of Fig. 6 we show also the integrated spectral weight 
N DDW (lom) and N(lom) at the same temperatures of the main panel, with a cut-off frequency uj^i = 6t. Even 
though this estimate of the spectral weight is much less accurate than the direct evaluation of Eq. (4.5), due to lowest 
numerical accuracy of this procedure, we find the same behavior discussed in the previous Sections while computing 



DDW 





19 



directly Eq. (4.5). Indeed, we can see that the spectral weight increases in the DDW state with respect to the normal 
state, even though the relative contribution of the Drude term is lower in the DDW state than in the normal state. 

A comment is in order now about the role of disorder. In the previous sections we reported the numerical results 
obtained for clean systems, but to amplify the differences between the conductivity of a non-interacting system and 
of the DDW state we used in Figs. 5, 6 a quite large value of the inverse scattering time w = O.lt, as appropriate 
for example to reproduce qualitatively the large Drude peak observed in BSCCO samples at about 100K [2, 4]. As 
a consequence, also the self-consistency equation (3.6) for the particle number and the sum rules W DDW (D, T) and 
W(T) should be evaluated in the presence of disorder for a given doping. The main difference is only in the absolute 
value of the spectral weight, while the relative difference between the case with or without DDW is the same. In 
the appendix B we discuss the modifications to the particle-number and spectral weight equations in the presence of 
disorder, that we used in computing the optical conductivity in Figs. 5, 6. 

V. DISCUSSION 

In the present paper we analyzed a possible approach to determine a GI approximation for the optical conductivity 
in a system which displays a transition to a d-wavc modulated CDW or flux phase. As we explained in detail in 
Sec. II the requirement of GI of a theory fixes the relation (2.20) between the current-current correlation function 
and the diamagnetic term. To understand better the expected domain of applicability the sum rule W DDW (D, T) 
(4.5) let us summarize the assumptions that led us to this rule. We have checked in Sec. IIIB (Eq. (3.13)) that the 
bubble II^ 7 ) of Eq. (2.21), evaluated with the mean-field DDW Green's function (3.4) and the bare vertex 7, does not 
satisfy the GI condition (2.20) when the standard diamagnetic term (3.7) is considered. As discussed in Sec. II C, this 
situation is quite standard, and considering the WI (2.26) this violation of the gauge invariance can be attributed to 
the k-dependent character of the DDW gap Z?k, which makes necessary the use of the full vertex T(fc, k) instead of 
the bare one 7(k, k). 

In general the vertex function is determined by solving the integral equation (2.24), but in the static limit it reduces 
to the expression (2.26). Since we do not know an analytical solution at finite frequencies and momenta of the vertex 
r(fc_|_,fc_), corresponding to the microscopic many-body Hamiltonian (2.1), we can try to use our knowledge of its 
static limit to give a better approximation than Eq. (2.21) for the current-current correlation function. More precisely, 
we showed in Sec. II D that the dc conductivity derived from the symmetric bubble Ilj| ra (2.33), where two full vertex 
functions in the static limit appear, coincides with the exact result at T = [39, 45]. Even though this procedure 
allows us to correctly reproduce the optical conductivity in the low- frequency limit, it does not solve the problem of 
knowing a priori the sum rule corresponding to this approximated optical conductivity. Indeed, since this symmetric 
bubble is not exact for the full model (2.1), and in contrast to the bubble (2.22) contains two full vertices, one cannot 
expect that the optical conductivity calculated using this bubble would satisfy the usual sum rule corresponding to 
the diamagnetic tensors (2.15) or (3.7). 

However, this last issue can be solved exactly by applying the same GI arguments to the reduced quadratic model 
(3.2). Indeed, the static limit Ti = % + Vi(k) of the full vertex function, obtained from the original interacting 
model, can also be considered as a bare vertex 7 for the DDW Hamiltonian (3.2). Moreover, since this Hamiltonian 
describes nonintcracting quasiparticlcs, the bare and the full vertex coincide, so that the symmetric correlator H DDW 
in Eq. (3.17) is the exact one for this model. As a consequence, the diamagnetic term (4.4) of the reduced model 
gives the sum rule W DDW (D,T) (4.5) for the symmetric current operator (3.17), which is the exact one within the 
quadratic theory (3.2) and at the same time provides us with a good approximation for the optical conductivity of 
the true interacting system, at least at low energy. 

The last issue we addressed in the present paper is to analyze to which extent the sum rule (4.5) can be related 
to the behavior of the microscopic Hamiltonian (2.1). In general, it is believed that in the presence of interactions 
the restricted sum rule (1.3), derived for the electrons within the lowest conducting tight-binding band, is still valid, 
provided that the occupation number n\^ a takes into account the effect of the interactions. In this case, we should 
rely on the estimate W(D,T) in Eq. (3.7) for the sum rule in the DDW state. However, this approach has two 
disadvantages: (i) we cannot derive the optical-conductivity which would lead to this sum rule; (ii) no general 
argument holds to justify why this attitude is the correct one to estimate, at mean-field level, the sum rule for the 
interacting microscopic model. Motived by these observations we argued that in the case of interactions leading to 
a DDW formation a better mean-field approach to the transport properties is provided by the calculation of the 
optical conductivity by means of the bubbles H DDW . Thus, to obtain the correct mean-field approximation for the 
spectral- weight behavior is not sufficient to modify the occupation number riko- below Todw, but it is more likely 
that a proper re-definition of the diamagnetic tensor is needed. As a consequence, the sum rule should be estimated 
by means of Eq. (4.5) instead of Eq. (3.7), leading to an increase of spectral weight below Tddw- However, this 
assumption would require also that the integrated spectral weight N DDW (u>) in the DDW state becomes larger than 



20 



the N(ci>) for the non-interacting system at some "low" frequency uim- As we discussed in Sec. IV D, such cut-off 
frequency u>m turns out to be in general lower than the plasma edge, but its precise value depends crucially on the 
parameters of the DDW transition (doping, order parameter at T — 0, etc.). Moreover, it is not clear yet if a u>m 
below the plasma frequency is a sufficiently low-energy scale for the interacting tight-binding model, since no universal 
definition exists of the frequency itself below which the restricted sum rule should be applicable. 

Finally, a comment is in order now about the comparison between our results and the experimental optical data for 
cupratcs. A first issue is related to the fact that the most recent experiments show that the temperature variation of the 
spectral weight is larger than expected in a tight-binding estimate also in optimally doped and overdoped compounds 
[4, 6] (see also Appendix A), i.e. eventually at doping larger than the critical doping 8$ = 0.2 for the charge ordering 
phenomenon. This could mean that a more general effect of the strong correlations present in these materials can 
be responsible for the large temperature variation of W(T). This possibility has been indeed investigated recently in 
Ref. [50], where W(T) has been evaluated by means of the DMFT (dynamical mean field theory) approach to the 
Hubbard model, which seems to reproduce the large temperature variations of W(T) observed in the experiments. A 
second issue arises about the lack, in the experiments, of a clear signature of an inter-band conductivity as the one 
reported in Fig. 4. In particular, BSCCO compounds, that were used as a paradigm for the choice of DDW parameter 
values in cuprates [26] , exhibit in general a quite featureless conductivity with a slowly decaying high-frequency tail. 
However, in different families of cuprates, displaying a similar spectral-weight behavior, clear signatures of charge 
ordering have been indeed observed in the optical spectra, even though located at much lower energy scales with 
respect to the one obtained here using the parameter values for BSCCO compounds. This is the case of LSCO and 
YBCO, where far-infrared features, well separated from the Drude peak, have been measured recently [47]. In both 
references [48, 49] these features were actually interpreted as due to a charge-ordering phenomenon, described by 
means of some different theoretical approaches which did not allow one to discuss at the same time the issue of the 
spectral-weight behavior. For these reasons, even though the analysis presented here cannot be conclusive as far as 
the optical spectra of HTSC are concerned, we believe that a deeper investigation of the role of charge degrees of 
freedom can eventually lead to a better understanding of the conductivity of cuprates. At the same time, the analysis 
presented here could be extended to other systems like 2if-TaSe2, where a fc-space modulated CDW forms [30] and 
where clear signatures of a Drude response accompanied by a mid-infrared peak have been observed in the optical 
spectra [29]. 

VI. ACKNOWLEDGMENTS 

We are grateful V.P. Gusynin for bringing to our attention the derivation of symmetrized expression for the deriva- 
tives of the polarization operator. We acknowledge stimulating discussions with A. Toschi, M. Ortolani and P. Calvani. 
This work was supported by the research projects 2000-067853.02/1, 620-62868.00, and MaNEP (9,10,18) of the Swiss 
National Science Foundation. 



APPENDIX A: FAILURE OF THE TIGHT-BINDING ESTIMATE 

As we noticed in the Introduction, even though a T 2 temperature decrease of W(T) is observed in the experiments, 
the measured slope is quite larger than the one expected within the simple non-interacting tight-binding estimate (1.4). 
To quantify this discrepancy in the most accurate way, we evaluate explicitly the spectral weight for the tight-binding 
model by including also a next-nearest neighbors term in the band dispersion, see Eq. (4.7). As a consequence, W(T) 
is given by Eq. (4.8). 

To correctly reproduce the Fermi surface of BSCCO and LSCO compounds we will assume t = 0.3eV and t' = rt, 
where r = 0.3 for BSCCO and r = 0.2 for LSCO (where the Fermi surface changes topology in the overdoped 
region, becoming electron-like at about 5 = 0.2 doping [51]). The results of W(T)/W(0) as a function of (T/t) 2 from 
Eq. (4.8) are reported in Fig. 7 for several doping (by fixing as usual the correct chemical potential at each doping 
and temperature from the self-consistency equation for the particle number) . In the left panel we report the estimate 
for BSCCO, that should be compared to the experimental data (for underdoped, optimally doped and overdoped 
samples) of Ref. [1, 2, 4]. Observe that in Ref.[2, 4] the variation of W(T) /W(0) between room temperature and 
T = of the order of 20% — 5% when measured at various cut-off frequencies, while the tight-binding estimate in Fig. 
7 never exceed the 0.6% (for t = 0.3 cV T = 300 K corresponds to (T/t) 2 = 0.0074). Analogous considerations hold 
for the comparison between the measured spectral weight in LSCO [6] and the estimate (4.8) reported in the right 
panel of Fig. 7. A comment is in order now about the role of the Van Hove singularity (VHS) in the density of states. 
Indeed, according to Eq. (1.4), where the t' = case was considered, the coefficient c(/i) = fiN'(fi) + N(fi) could 
increase considerably by approaching the VHS. This effect is indeed seen in the curves at 6 = 0.12 and 6 = 0.16, where 



21 




FIG. 7: (Color online) Spectral weight W(T)/W(0) according to Eq. (4.8) for t' = 0.3t (left panel, appropriate for BSCCO 
[2, 4]) and t' = 0.2t (right panel, appropriate for LSCO [6]) at various doping. In the right panel we also report the result 
obtained with the estimate (1.4) and the parabolic band (PB in the caption) approximation for c(/x) = l/47rf, using the value 
W(0) = 0.6824 at 6 = 0.26. 



the initial slope of W(T) is quite large. However, as soon as the temperature increases the effect of the VHS is washed 
out very rapidly and the overall variation in the range of T between — O.li attains the same values found for the case 
t' = 0.3i. Moreover, for the overdoped case 8 = 0.26 the slope of W(T)/W(0) agrees very well with the approximation 
c(/n) = 1/Airt of the parabolic band dispersion (PB in the figure), which would give the value 7r 2 c(/z)/6 = 0.13/t for 
the coefficient in Eq. (1.4). For these reasons one can conclude that the t' term in the band dispersion (4.7) has a 
minor role in determining the spectral-weight variations, and indeed it was only briefly discussed in Sec. IV C of the 
present work. 



APPENDIX B: SUM RULE IN THE PRESENCE OF DISORDER 



As we did in Sec. IV D we will take into account the effect of disorder by using the substitution (4.11) in the spectral 
representation (3.10) of the Green's function. To see how Eq. (3.6) is modified we rewrites it in terms if the spectral 
function: 

2^ RBZ „ RBZ 



n = — 
N 

k 



Tr[GQc, iu n )]e^ 0+ = A £ / dz[M(z - E) + M(z + E)]f(z - M ). (Bl) 



Analogously, the definition (4.4) of the diamagnetic tensor in the DDW state can be expressed as: 

rp RBZ 

(tu) - -jj {£kTr[G(k,io; n V3]-£> k Tr[G(k,ia; ri )cr 2 ]}e^ 0+ = 



. RBZ 



Observe that if one puts M(z) = S(z) the results (3.6) and (4.5) can be recovered, and for Do = one finds the 
corresponding expressions for the normal state. At T=0, which is the case considered in Fig. 5, the previous equations 
simplify. Indeed, since the Fermi functions reduce to a step function, one has: 

1 1 { i E tt\ 
dz „ W 9 = - ( arctan — h - ) (B3) 

.oo 7T Z Z + W Z 7T \ W 2 J 



22 



so that the self-consistency equation for the particle number and the equation for the spectral weight can be written 
as: 

2 f fx - E fx + E 



. 6 \ ■> / U — Hi U, -T Hi \ /tims 

n — 1 = > arctan h arctan , (B4) 

nN ^-^ V w in I 

k 

RBZ 



W DDW (D,T = 0) l ^ f fi + E\ 

; — „ v ,' ; = > E k [ arctan - arctan - . (B5) 

(-ne 2 a?/V) irN ^ k V w w J V ' 

Observe that here we did not consider the effect of the DDW formation on the transport scattering time, which can 
be present. For a detailed discussion of impurity scattering in the DDW state see Ref. [35]. 



[1] H.J.A. Molegraaf, C. Presura,D. van der Marel, P.H. Kes, M. Li, Science 295, 2239 (2002). 

[2] A.F. Santander-Syro, R.P.S.M. Lobo, N. Bontemps, Z. Konstantinovic, Z.Z. Li, H. Raffy, Phys. Rev. Lett. 88, 097005 

(2002); Europhys. Lett. 62, 568 (2003). 
[3] C.C. Homes, S.V. Dordevic, D.A. Bonn, Ruixing Liang, W.N. Hardy, Phys. Rev. B 69, 024514 (2004). 
[4] A.F. Santander-Syro, R.P.S.M. Lobo, N. Bontemps, cond-mat/0404290. 

[5] A. V. Boris, N. N. Kovaleva, O. V. Dolgov, T. Holden, C. T. Lin, B. Keimer, C. Bernhard, Science 304, 708 (2004). 
[6] M. Ortolani, P. Calvani, S. Lupi, cond-mat/0407432. 
[7] J. Jaklic and P. Prelovsek, Adv. Phys. 49, 1 (2000). 

[8] A.J. Millis, Strong Interactions in Low Dimensions, ed. by D. Baeriswyl and L. De Giorgi, Kluver Accad. Pub, 2003. 
[9] D. van der Marel, in "Interaction Electrons in Low Dimensions", book series "Physics and Chemistry of Materials with 
Low-Dimensional Structures", Kluwer Academic Publishers (2003); cond-mat/0301506. 
[10] PF. Maldague, Phys. Rev. B 16, 2437 (1977). 

[11] D. Baeriswyl, C. Gros, and T.M. Rice, Phys. Rev. B 35, 8391 (1987). 

[12] D.J. Scalapino, SR. White, and S. Zhang, Phys. Rev. B 47, 7995 (1993). 

[13] See L. B. Ioffe and A. J. Millis, Science 285, 124 (1999) and references therein. 

[14] J.E. Hirsch, and F. Marsiglio, Phys. Rev. B 62, 15131 (2000). 

[15] MR. Norman, and C. Pepin, Phys. Rev. B 66, 100506 (2002). 

[16] T. Eckl, W. Hanke, and E. Arrigoni, Phys. Rev. B 68, 014505 (2003). 

[17] A. Knigavko, J. P. Carbotte, and F. Marsiglio, cond-mat/0404153. 

[18] V.M. Loktev, R.M. Quick, S.G. Sharapov, Phys. Rep. 349, 1 (2001). 

[19] B.I. Halperin, and T.M. Rice, Sol. State Phys. 21, 115 (1968). 

[20] I. Affleck, J.B. Marston, Phys. Rev. B 37, 3774 (1988). 

[21] A. A. Nersesyan and G.E. Vachanadze, J. Low Temp. Phys. 77, 293 (1989). 

[22] H.J. Schulz, Phys. Rev. B 39, 2940 (1989). 

[23] CM. Varma, Phys. Rev. B 55, 14554 (1997). 

[24] M.V. Eremin, LA. Larionov, Pis'ma Zh. Eksp. Teor. Fiz. 68, 583 (1998) [JETP Lett. 68, 611 (1998)]. 

[25] E. Cappelluti and R. Zehyer, Phys. Rev. B 59, 6475 (1999). 

[26] L. Benfatto, S. Caprara, and C. Di Castro, Eur. Phys. J. B 17, 95 (2000). 

[27] S. Chakravarty, R.B. Laughlin,D.K. Morr, and C. Nayak, Phys. Rev. B 63, 094503 (2001). 

[28] C. Castellani, C. Di Castro, and M. Grilli, Phys. Rev. Lett. 75, 4650 (1995). C. Di Castro, L. Benfatto, S. Caprara, C. 

Castellani, M. Grilli, Physica C 341-348, 1715 (2000). 
[29] V. Vescoli, L. Degiorgi, H. Berger and L. Forro, Phys. Rev. Lett. 81, 453 (1998). 
[30] A. H. Castro Neto, Phys. Rev. Lett. 86, 4382 (2001). 
[31] L. Benfatto, S. Sharapov, and H. Beck, Eur. Phys. J. B 39, 469 (2004). 
[32] Y. Nambu, Phys. Rev. 117, 648 (1960). 

[33] J. R. Schrieffer, Theory of superconductivity, Addison Wesley (1988). 

[34] Q.-H. Wang, J.H. Han, and D.-H. Lee, Phys. Rev. Lett. 87, 077004 (2001). 

[35] J.-X. Zhu, W. Kim, C. S. Ting and J. P. Carbotte, Phys. Rev. Lett. 87, 197001 (2001). 

[36] H.-Y. Kee and Y.B. Kim, Phys. Rev. B 66, 052504 (2002). 

[37] X. Yang and C. Nayak, Phys. Rev. B 65, 064523 (2002). 

[38] S. Chakravarty, C. Nayak, S. Tewari, and X. Yang, Phys. Rev. Lett 89, 277003 (2002). 
[39] S.G. Sharapov, V.P. Gusynin, and H. Beck, Phys. Rev. B 67, 144509 (2003). 
[40] D.N. Aristov and R. Zeyher, cond-mat/0405194; cond-mat/0406419. 
[41] B. Valenzuela, E.J. Nicol, and J. P. Carbotte, cond-mat/0407497. 

[42] CP. Enz, A Course on Many-Body Theory Applied to Solid-State Physics, (World Scientific, Singapore, 1992). 

[43] The definition (2.16) for a lattice system is correct only in the limit of small q, when one can approximate £k+ q — £k ~ q-^k • 
However, since in the following we are interested in discussing the optical conductivity, which is related to the q = limit 
of the current operator (2.14), we can adopt the notation (2.14)-(2.16) without loss of generality. 



23 



[44] V.P. Gusynin, V.A. Miransky, Mod. Phys. Lett. A 6, 2443 (1991); Sov. Phys. JETP 74, 216 (1992). 
[45] J.S. Langer, Phys. Rev. 127, 5 (1962). 

[46] The Hamiltonian (3.2) does not display any explicit charge modulation, even though a gap appears in the quasiparticle 
spectrum. To describe a true CDW modulation a coupling like |-Dk|c£ CT Ck+Qo- + h.c. should be considered instead. 

[47] A. Lucarelli, S. Lupi, M. Ortolani, P. Calvani, P. Maselli, M. Capizzi, P. Giura, H. Eisaki, N. Kikugawa, T. Fujita, 
M. Fujita, and K. Yamada Phys. Rev. Lett. 90, 037002 (2003); C. Bernhard, T. Holden, A. V. Boris, N. N. Kovaleva, 
A. V. Pimenov, J. Humlicek, C. Ulrich, C. T. Lin, and J. L. Tallon, Phys. Rev. B 69, 052502 (2004). 

[48] S. Caprara, C. Di Castro, S. Fratini, and M. Grilli, Phys. Rev. Lett. 88, 147001 (2002). 

[49] L. Benfatto and C. Morais Smith, Phys. Rev. B 68, 184513 (2003). 

[50] A. Toschi, M. Capone, C. Castellani, M. Ortolani and P. Calvani, unpublished. 

[51] A. Ino, C. Kim, M. Nakamura, T. Yoshida, T. Mizokawa, A. Fujimori, Z.-X. Shen, T. Kakeshita, H. Eisaki, and S. Uchida, 
Phys. Rev. B 65, 094504 (2002).